You are on page 1of 37

Journal Pre-proof

Exergy analysis and assessment of a new integrated industrial based energy system
for power, steam and ammonia production

Ghassan Chehade, Ibrahim Dincer

PII: S0360-5442(19)31972-3
DOI: https://doi.org/10.1016/j.energy.2019.116277
Reference: EGY 116277

To appear in: Energy

Received Date: 2 August 2018


Revised Date: 20 August 2019
Accepted Date: 2 October 2019

Please cite this article as: Chehade G, Dincer I, Exergy analysis and assessment of a new integrated
industrial based energy system for power, steam and ammonia production, Energy (2019), doi: https://
doi.org/10.1016/j.energy.2019.116277.

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. Please note that,
during the production process, errors may be discovered which could affect the content, and all legal
disclaimers that apply to the journal pertain.

© 2019 Published by Elsevier Ltd.


Exergy Analysis and Assessment of a new Integrated Industrial Based Energy System for
Power, Steam and Ammonia Production

Ghassan Chehade and Ibrahim Dincer


Clean Energy Research Laboratory, Faculty of Engineering and Applied Science, University of
Ontario Institute of Technology, 2000 Simcoe Street North, Oshawa, Ontario L1H 7K4, Canada
Email: Ghassan.Chehade@uoit.ca, Ibrahim.Dincer@uoit.ca

Abstract
In this study, a new integrated energy system for power, steam and ammonia production is
developed, simulated through Aspen Plus and assessed thermodynamically through energy and
exergy approaches. The system employs a novel ammonia production loop to potentially replace
the conventional Haber-Bosch process by integrating an expander/turbine in the ammonia loop.
The ammonia production system produces 2310 kmol/h of liquid ammonia and steam as a by-
product. Although having an optimum conversion rate is attained by controlling/shifting the
volume of synthesized ammonia, the exergetic assessment shows that the best output ammonia
product is achieved by the performance of each independent cycle. Hence, the overall exergy
efficiency of the system is 71 % and turbine exergy efficiency of 92 % for optimal operating
conditions with a total work rate of 12631 kW and ammonia conversion rate of 38%
respectively.
Keywords: Energy, exergy efficiency, ammonia synthesis, steam reforming, methanation, heat
recovery.

1 Introduction
It is evident that ammonia synthesis from the elements; hydrogen and nitrogen, is amid among
the greatest innovation of the twentieth century. Ammonia provides most of the fertilizer that we
use today and the importance cannot be exaggerated as the development in the world population
and our ability to produce ammonia are closely correlated [1]. It is important to note that the
ammonia industry is well established among the high energy-consuming process, which has been
modernized in the last few years by heat and cold recovery approaches. However, exergy
analysis studies have proved that the transitional process of steam generation in the reformer,
ammonia synthesis loop, and water gas shift is not efficient [2]. Thus, integrated multigeneration
platform (heat and power combined with chemical energy) evoke synergistic effects which can
improve the overall system efficiency and reduce greenhouse gas intensity of the end-use sectors.
1
The application of the second law of thermodynamics in the ammonia industry especially in the
synthesis loop and heat recovery, presents an attractive opportunity for possible improvements.
Greef et al. [3] investigated the direct integration of a gas turbine in the ammonia synthesis loop
using energy and exergy analysis. They found that the pressure ratio across the turbine plays an
important role in determining the conversion rate in the ammonia synthesis reactor. Also, the
results show a decrease in exergy loss when compared to the conventional ammonia plant. Florez
et al. [4] performed a study on the industrial ammonia unit based on steam methane reforming.
The system had a capacity of 1000 tonne NH3/day. The results, showed that the maximum
converion of synthesis gas is attained by controlling the inlet temperature of the reactor beds
using waste heat boilers. Also, the exergy destruction across the ammonia synthesis loop
accounted for more than 82%. In another study also performed by Florez et al. [5] for a 1000
tonne NH3/day industrial ammonia unit, three synthesis loop configurations were proposed and
analyzed using exergy efficiency. The results showed that the maximum irreversibilities come
from the compressor, ammonia converter, and refrigeration system which accounts for more than
86%. Kirova [2] analyzed the ammonia synthesis loop in an exergy study. The study showed the
ammonia synthesis reaction and the heat exchanger are the major sources for exergy losses
accounting for more than 65% and the only way to decrease the exergy loss is by utilizing the
reaction temperature through heat recovery in multiple phases. Khademi et al. [6] simulated and
optimized the ammonia synthesis loop by implementing cooling methods in three different
scenarios: internal direct cooling reactor (IDCR), adiabatic quench cooling reactor (AQCR), and
adiabatic indirect cooling reactor (AICR). The main objective of the optimization study was the
maximum N2 conversion. The results showed that the maximum conversion of 0.26 was
achieved in the 3-bed AQCR at an inlet temperature of 635 K at the first bed. However, the
IDCR achieved a conversion ratio of 0.3 and a gas temperature of 495 K. While, AICR achieved
a conversion ratio of 0.3 in the 3-bed at inlet gas temperature of 696 K for each bed. Sahafzadeh
et al. [7] performed a study to investigate and improve the energy efficiency of the ammonia
synthesis process. The ammonia synthesis loop was integrated with a gas turbine to reduce the
exergy loss associated in the ammonia converter. The results showed that the gas turbine
accounted for 4 MW of generated work at a penalty of adding 7350 kW of high-pressure steam.
Moreover, an integrated energy system for ammonia synthesis and power generation was
proposed and analyzed by Aziz et al. [8]. The combined system adopted the Haber-Bosch
2
process in the ammonia synthesis loop as well as a power generation module to recover heat in
the ammonia converter and supply it to the nitrogen production unit. The overall energy
efficiency was about 66.92% and the NH3 conversion ratio was found to be 0.63. Rosen [9] did a
thermodynamic assessment based on energy and exergy analysis of an SMR process for
hydrogen production from natural gas. A numerical code was developed to identify state
properties and system characteristics such as exergy losses and destruction through the proposed
layout. The results revealed that the main exergy losses took place in the reformer and were due
to the irreversibilities associated with the combustion process and the places where heat transfer
occurred at greater temperatures. Furthermore, Simpson et al. [10] analysed steam methane
reforming energetically and exergetically. They concluded that although exhaust gases entail a
significant amount of exergy, a considerable amount of exergy is destroyed due to irreversible
heat losses and chemical reactions.
Although studies were conducted on the ammonia synthesis system and methodologies
on improving the energy and exergy efficiencies, efforts have not been exerted toward the
development and investigation of a comprehensive modern ammonia production system (steam
methane reforming, water-gas shift reactor, carbon dioxide removal, methanation, ammonia
converter, heat recovery and power generation). Here, an integrated ammonia production system
is newly developed by combining both a classical and innovative approach for practical
applications. Furthermore, the performance of the integrated ammonia production unit is
assessed using exergy efficiency. The specific objectives of the study are (i) to develop a novel
integrated process for the ammonia production system, (ii) to model and simulate the developed
system in the Aspen Plus simulation software, (iii) to assess the system performance in terms of
exergy efficiency.

2 System description
The reaction of hydrogen and nitrogen is driven by three main properties: temperature, pressure
and volume. Thus, the proposed system characteristics are developed based on the law of Le
Chatelier. The overall schematic diagram is shown in Fig. 1 consists of five subsystems, the
reformer unit which consists of primary and secondary reforming, the CO conversion unit which
consists of a high-temperature-shift reactor and a low temperature-shift reactor, the CO2 removal
unit, methanation, dryer, the ammonia synthesizer, and the integrated 2 stage expander unit.
3
Methane is mixed with superheated steam and is feed to the primary reformer; the steam-gas
mixture is then heated further between 500 and 600oC since the process of reforming is highly
endothermic, additional heat is required, so the temperature is raised to up to 830 oC. The nickel-
based reforming catalyst has no effect on the position of chemical equilibrium instead it provides
an alternative pathway with lower activation energy and hence increase the reaction rate.
However, 30 % to 40 % of the hydrocarbons react in the primary reformer. Thus the products are
blended with air before flowing into the secondary reformer for an additional yield of hydrogen.
In the high-temperature shift conversion process, undesired carbon monoxide is removed; the gas
is passed on a bed of iron oxide and chromium oxide catalyst at a temperature of 400 oC and is
converted to carbon dioxide (CO2) which is then removed by absorption in the aqueous solvent.
Withal, a small amount of carbon monoxide and carbon dioxide remaining in the process are
converted to methane in the methanation process. The reaction takes place at around 300 oC.
Finally, nitrogen and hydrogen mixture pass to the ammonia synthesis unit. In the ammonia
synthesizer, the chemical reaction is highly exothermic; it releases energy so that the sum of the
enthalpies of N2 and H2 (the reactants) is higher than the enthalpy of NH3. The reaction is
affected by temperature, and pressure on the composition of the equilibrium mixture, the rate of
the reaction and the economics of the process. The synthesis of ammonia takes place in an iron-
based catalyst at high pressure and temperature of 246.7 atm and 550 ◦C respectively, only 20 to
30% of hydrogen and ammonia are reacted due to the unfavorable equilibrium conditions. Since
the reaction of hydrogen and nitrogen is considered highly exothermic and reversible the need to
find more efficient ways to extract ammonia from the synthesis loop is essential. In the
integrated ammonia loop, the reaction heat of the exothermic reaction is used to produce power
in the expander. The hot reaction products are expanded directly thus resulting in low exergy loss
and consumption and better conversion rate across the synthesis loop. Furthermore, the power
generated by the turbine/expander is used to power the compressors and lower the energy
consumption across the system. Finally, unreacted gas of hydrogen and nitrogen and produced
ammonia are separated by a liquid-vapor separator. The extracted ammonia can be pumped to a
storage tank while the excess hydrogen and nitrogen are compressed and then introduced back to
the synthesis reactor as illustrated in Fig. 1.

4
3 System modeling
The performance of the ammonia production unit is modeled using exergy method. Mass,
energy, entropy, and exergy balance equations are defined for all the subprocess and are
implemented in Aspen Plus software. To provide accurate results for the thermodynamic
properties of the high temperature, high-pressure conditions, Redlich-Kwong modification RKS-
BM is selected. Henry-components are selected for the compound NH3. Also, RK equation of
state and Electrolyte NRTL method is used to calculate both liquid and vapor properties in the
carbon dioxide removal subsystem. The overall system design is presented in Fig. 2 with
temperature and pressure shown at the main inlets and outlets of the subprocesses. Also, room
temperature (298 K) and atmospheric pressure (101.325 kPa) are taken as the reference point.
While all the compressors and turbines are assumed to have 72 % isentropic efficiency and 100
% mechanical efficiency.
3.1 Exergy balance equations
The exergy balance equation of a general system disregarding any potential or kinetic change is
expressed as follows:
∆ = ∆,
 + ∆,   (1)

  +   1 −  + ! " +  # +    


 !


=   +  $ 1 −  + %
 ' +  #$ +     +   ( )*+ ) ,- (2)
 &

The chemical exergy based per mole of the mixture in Table 2 is expressed as follows:
///    + 2  01 ln 01 " (3)
   =  01 ///
1

For hydrocarbons the chemical exergy is expressed regarding Gibbs functions as follows [11]:
J
HI L
:
G> " K
///
   = 6 7̅9 + % + ;
<' 7̅=> − 7̅?=> − @ 7̅A> = B + 2 ln F L M (4)
(CD ,ED ) (?=> )H (A> =)K

3.2 Reactor kinetic modeling for the reformer unit


In the reformer unit, methane and steam are converted to a mixture of H2, CO2, and CO. Figure 3
represents the flowsheet of the reformer unit (subsystem A). SREF-R and PREF-T are modeled
as Rplug reactors and represent the number of firing zones. Also, PREF-S and SREF-S are
RSTOIC reactors and represent the burners in which combustion is complete. Steam mixed with
methane enters the reformer unit at 502.55 ◦C and 3530.39 kPa. The reactors are connected in
series and have a total number of 560 tubes. Each tube is 10 m long and has a nominal diameter
5
of 0.14 m. The reaction in the reformer unit is highly endothermic thus, burning fuel is required
to achieve the reaction. Air is fed into the reformer unit at a flowrate of 0.66 kmol/s and
temperature and pressure of 492 ◦C and 3236.19 kPa respectively. This is achieved by 4 stage
compression with inter cooling. The oxygen in the air reacts and raises the temperature to 1200

C so the reactions take place:
a. CH4 + H2O → CO + 3H2
b. CO + H2O → CO2 + H2
To accurately model the rate of reaction in the tubes, the heat transfer rate is calculated by the
flame temperature (975 ◦C), gas temperature (1200 ◦C) and the heat transfer coefficient of the
tube (0.001163 kW/m2). The modeling of the reformer subsystem is based on the published work
of Alatiqi et al [12]. The reformer chemical kinetic expression is as follow:
O@ PQ (X@ ); (WY@ ) @
2= UV (WX )( X Y ) @
− P Z (5)
R Q 379 @ ; @
R@
The reaction rate with reference to the environment is expressed as follows:
]J^>D
O@ = \  (6)
% ab.d:@'
_IK`D

Also, V 2 = V@ × VQ then the terms follow


;d,;Qj
V: = exp %− Ck;l + 30.707' no@ (7)
p,@;
V@ = exp % − 4.335' qrst 1,100 ℉
Ck;l

7,351.24
VQ = exp  − 3.765 nqsv 1,100 ℉
 + 460
The chemical equilibrium is achieved when:
(X@ )(WY@ )
VQ = (8)
(X@ Y)(WY)
3.3 Reactor kinetic modeling for the CO-shift unit
The product stream leaving the reformer unit is cooled down via the heat exchanger HEX2 to a
temperature of 380 ◦C. Energy is recovered in HEX2 and is feed to the power generation unit.
The cooled stream S3 goes into the carbon monoxide removal unit (subsystem B). The unit
consists of 2 reactors in series: High-temperature shift reactor (HT-SHIFT) and the low-
temperature shift reactor (LT-SHIFT), as presented in Figure 6. The HT-SHIFT is modeled as
RPlug reactor and is used to convert some of the carbon monoxides and recover some of the heat
using the heat exchanger HEX-9 which is fed back to the power generation unit. Then the stream
6
CO-10 goes into the LT-SHIFT. The reaction rate in the LT_SHIFT is low due to the low
operating temperature of the reactor but attains higher conversion rate. The kinetic expression for
the low-temperature shift is described as follows [13]:
:
{ V9 ‚
*$ O|C }?= }A> = %1 − VQ '
@

x?= = \ z  (9)
 z1+O } +O } 
yP
~ ?=  ?=>
€
The high-temperature conversion plug reactor is modeled as follows:

V$
 (10)
:
2?= = \ VAC P@ }?= 1 −
VQ
3.4 Electrolyte modeling for the CO2 removal unit
The carbon dioxide is removed by absorption in the aqueous solution. The electrolyte solution
chemistry is modeled in Aspen Plus, where chemical equilibrium is assumed with all the ionic
reactions in salts. A lean aqueous ammonia solvent (CO2-R1 stream) is fed to the top of the
absorber at a flowrate of 3.014 kmol/s with a composition of 15 mol% ammonia (NH3) and 85
mol% water as shown in Figure 8. The carbon dioxide is entirely absorbed by the solvent. The
gas leaving the absorber contains 0.002 kmol/s of CO. The mixture leaves at the base of the
absorber is then regenerated in the stripper by raising the temperature to 60 °C and lowering the
pressure to atmospheric. Various chemical reactions associated with the aqueous ammonia-based
CO2 absorption process that are simulated by ASPEN Plus are given below:
i. ƒXQ + X@ Y ↔ ƒX;k + YX a Equilibrium

ii. 2 X@ Y ↔ XQ Yk + YX a Equilibrium

iii. XWYQa + X@ Y ↔ WYQa@ + XQ Yk Equilibrium

iv. WY@ + YX ↔ XWYQa Kinetic

v. XWYQa ↔ WY@ + YX a Kinetic

vi. ƒXQ + WY@ + X@ Y ↔ ƒX@ WYY a + XQ Yk Kinetic

vii. ƒX@ WYYa + XQ Yk ↔ ƒXQ + WY@ + X@ Y Kinetic

viii. ƒX; WYQ ( ) ↔ ƒX;k + XWYQa Electrolyte

ix. WY@ + 2ƒXQ (n…) → (ƒX; )@ WYQ Electrolyte

The reaction rate of the CO2 removal is modeled by the power-law expression as follows [14]:

7
‹
 - − 1 1
x = O   ‡ ˆ   − ‰ Š W ! (11)

 2  
Œ:

x
‹

= O exp −  Š W ! (12)
- 
2
Œ:

The values for K and E are given in Table 3 while the molarity n is zero.
The built-in Keq expression is used for the salt precipitation reaction of NH4HCO3:
Ž
ln V = \ + + W ln() + 

The values for A, B, and C are presented in Table 4.
3.5 Reactor chemical kinetics for the Methanation unit
In the methanation unit, excess carbon monoxide and carbon dioxide react to form hydrogen in
the presence of the nickel-based catalyst Ni/Al2O3. The reactor RM-1 in subsystem D
represented in Figure 10 is modeled as Rplug. The reactions associated in the methanation
reactor:
x. CO +3H2 →CH4 + H2O
xi. CO2+4H2→CH4 + 2H2O
The reaction rate is based on the Langmuir-Hinshelwood mechanism as follows [15]:
: : P <?AK <A@>=
x = \ 0.314 %:,Q6Caj:QB'  ‘ U< − Z (13)
PA.j
>
?=>
<A;> P@ V?=>

The reaction rate for CO in methanation is expressed as follows:


@:,l@:
V?=> =  %aQQ.d@Qk C
'
(14)

3.6 Chemical equilibrium and Kinetics for the Ammonia synthesis unit

The gas mixture entering the ammonia synthesis unit (subsystem E) is composed mainly of
hydrogen and nitrogen and small quantities of argon as presented in Figure 11. Also, the
unreacted gas recycle stream (S15 or SYN-2) is mixed with the gas mixture entering the first bed
(R1-A) and the total gas mixture is compressed to 30000 kPa. The stream SYN-8 leaving the
first bed has a temperature of 532 °C. The exit stream of the first bed (SYN-8) is cooled down
with cooler gas in HEX 10 and then enter the second bed (R1-B) at an inlet temperature of 456
8

C. The inlet stream SYN-11 is then cooled down by the heat exchanger HEX11 before entering
the final bed R1-C. The inlet temperature of the third bed is 395 and the outlet temperature
reaches 431 ◦C before entering the Converter R1-D. The 3 reactor beds have a length of 2.4m and
a diameter of 2m. The reaction taking place in the ammonia synthesizer is exothermic. The
partial mole of the gas mixture of nitrogen and hydrogen in terms of reaction conversion ξ is
expressed as follows:
n’ @, − 0‹>,& n“ @, − 0A>,& n’“ Q, − 0‹A],&
ξ = = = (15)
0‹>,! 0A>,! 0‹A],!

The reaction rate of ammonia synthesis is expressed using the Temkin-Pyzhev equation in an
iron catalyst bed as follows [16][17]:
.jj .;j
•AQ •‹A
@
Osr
x‹A] = 2O” F V@•‹> – @ > — −– ]
— M U Z (16)
•‹A] •AQ>  )
)
Q
.ℎ

In this case, the fugacity, • at equilibrium is related to the fugacity of the pure component •  at
the same pressure and temperature follows Lewis-Randall rule [17]:
• =  × •  (17)
The pure component fugacity as mentioned is found at reactor pressure P and temperature:
•  = ™ × P (18)
Also, the equilibrium constant is found using the expression:
O$
V@ = (19)

The pressure drop across the ammonia reactor is modeled using Ergun correlation [18]:
¤›œ ¤›œ
(E ›œ :a  :j:a ¥ ›œ
= − ž × ×F + 1.75 M
¤Ÿ œ¡¢£Ÿ ¤Ÿ œ¡¢£Ÿ
]
(š ¦§ ~
¤›œ
(20)
Ÿ œ¡¢£Ÿ  
¤Ÿ œ¡¢£Ÿ

3.7 Exergy analysis

Exergy has the characteristic that it is not conserved only when all processes occurring in a
system and the environment are irreversible. When an exergy analysis is performed on a
chemical processing plant, the thermodynamic imperfections can be quantified as exergy loss
and consumption, which represent losses in energy quality or usefulness [19]. In this study,
exergy efficiencies are evaluated based on two characteristics, the first exergy efficiency is based
on the useful exergy products of the system compared to that of the exergy input of the system.

9
However, this takes only into account exergy transformed by the process of synthesizing
ammonia and is presented in equation (I). The second exergy definition (II) is used to access the
quality of the steam recovered in the process. Since the amount of steam generated is much
higher than the produced ammonia, steam recovery becomes important, and the quality of steam
shows the irreversibilities in the subprocess where steam can be recovered. Unreacted gases that
are in flow or transit, such as methane, argon and excess gas are purged out of the system. The
amount of purge affects directly the quality of ammonia produced. Equation (III) is proposed to
access the quality of purged gas. If the flow of the purged gas is high the exergy efficiency also
will be high and sometimes misleading. Thus, accounting for the purged and transit gas flow it is
important to take into consideration the irreversibilities due to purging. Considering all three
exergy efficiency definitions, the first exergy efficiency definition gives a better understanding
of the performance of the overall system. Also, while assuming all the excess gas is converted
into ammonia. Equation I, II and III are derived on the consideration that the work generated by
the turbine expander and the produced ammonia in line with the expander. So, the approach
suggests, recovering work in the form of mechanical work to drive the pumps, compressors and
extract heat of produced ammonia directly after synthesis reactor results in decrease total power
consumption and increase production rate could be considered as a more attractive way and thus
the exergy efficiency of the system is described in terms of expander work and ammonia
recovered to that of steam recovered in the process. Although equation (II) reflects the potential
for exergy recovery, the main purpose of the system is to produce ammonia. Thus, deviating
from comparing the low- quality exergy of steam being produced, equation (I) is derived based
on the ammonia reactor minimum exergy consumption to the actual consumption of natural gas
and steam going into the system and generated work.
4 Results and discussion

The performance of the reformer unit/subsystem A is evaluated based on exergy loss and
consumption, and exergy efficiency. Table 1 illustrates the operating base parameters for the
integrated system. In the reformer subsystem as shown in Fig. 3, hydrogen is produced at high
temperature and pressure. The amount of hydrogen produced is found to be 0.82 kmol/s. Taking
into consideration that the 3- plug reactors PREF-S, PREF_T and SPREF-R represent the firing
zone of a reformer unit and graphing the enthalpy flow versus temperature as shown in Fig. 5,
the reaction in highly endothermic with an enthalpy reaction rate of 35000 kW. The carbon to
10
hydrogen ratio is calculated at 0.81 while, the overall exergy efficiency of the reformer unit is
78.16 % as presented in Fig. 15.
The reformer exergy results are compared with the literature for validation. Srin et al.
[20] found an exergy efficiency of 79.88 % of a reformer unit where heat is provided by
combustion. Also, Rosen [9] reported an exergy efficiency of 78.5 % for a heat integrated steam
reforming unit. While Bargigli et al. [21] performed a comparative study to access hydrogen
production and reported an exergy efficiency of 71 %. The products of the reformer unit are
cooled in the waste heat boiler before entering the high-temperature water shift reactor as shown
in Fig. 6.
In the high-temperature shift conversion unit in subsystem B, carbon monoxide reacts
with steam to produce hydrogen and carbon dioxide. The reaction is exothermic and depends on
the carbon monoxide concentration. The exergy analysis results in subsystem B are illustrated in
Table 6. The high-temperature shift exergy loss and consumption are found to be 2335 kW while
the low-temperature shift reactor unit has an exergy rate of 743 kW as presented in Fig 7. The
drastic change in the exergy consumption and loss is due to the change in temperature across
both reactors and a decrease in the concentration of CO. The cooling process is achieved by
integrating a boiler feed water heat exchanger and the by-product is superheated steam exiting
the system at a temperature of 300 °C. The overall exergy efficiency of the CO shift conversion
unit is 93.63% which is considered moderate when compared to the reactor unit proposed by
Rosen [9].
In the CO2 removal unit/ subsystem C as shown in Fig. 8, the absorber reactor absorbs
most of the CO2 in the presence of aqueous NH3, and the by-products are hydrogen carbonate as
shown in Table 7. The implementation of ammonia solvent is considered new in the chemical
industry, and the reason is that ammonia solvent solves degradation problems and increase the
capacity of capturing CO2. Also, unlike alkanolamines ammonia solvent has lower heat of
absorption [22] and thus making the process less energy intensive.
Fig. 9 shows the exergy losses and consumptions across the carbon dioxide removal unit.
Also, the overall exergy efficiency of the CO2 removal unit/subsystem C is found to be 67.75 %.
The remaining CO and CO2 in stream CO2-R3 presented in Table 8 are then converted into
methane via the methanation process/subsystem D as presented in Fig. 10. The exergy efficiency
of the methanation unit is 97% and exergy consumption and loss is 6000 kW. The product of the
methanation unit is called synthesis gas that is made of hydrogen and nitrogen. The
11
stoichiometric ratio is adjusted before the synthesis unit to achieve 3H2 for 1 N2 mole. After the
methanation process, the synthesis gas is compressed to 30000 kPa as presented in Table 9 and
H2 and N2 are catalytically converted into NH3 in the synthesis unit/ subsystem E as shown in
Fig. 11. The exergy efficiency of the ammonia unit is 98% and the conversion fraction is found
to be 0.38. The reaction between hydrogen and nitrogen is exothermic and this is noted from the
increase in the enthalpy of formation across the synthesis unit/subsystem E as presented in Fig.
12.
In order to extract the volume of produced ammonia and utilize heat rejection a 2-stage
expander unit is integrated into the ammonia loop subsystem to produce power. The work
produced by the expander is used to operate the compressors C-1 and C-2 and C-3. The ammonia
produced is then separated in a separator based of phase change with an achieved production rate
of 2232 kmol/h and the unreacted gases are then compressed to 30000 kPa, then heated up to 180
°C and are recycled back into the synthesis unit. The exergy analysis result for the ammonia loop
is presented in Table 10. Also, Fig. 13 shows the conversion ratio of ammonia across R1-A, R1-
B, and R1-C catalyst bed and is compared to the optimal equilibrium conversion ratio of 0.19 at
400°C.
The net power generated by the integrated system is found to be 12631 kW and this
includes the power generated by the steam turbine in the power generation subsystem and the
power consumed by the compressors across the system. Based on the proposed integrated system
for power steam and ammonia production and with the thermodynamic properties it is feasible to
access the system performance through exergy methods. The highest exergy loss and
consumption rate is attained by the carbon dioxide removal unit while the lowest exergy loss and
consumption rate is found in the methanation process as illustrated in Fig. 14.
Furthermore, Fig. 15 shows the exergy efficiency of the proposed subsystems. The heat
consumption and rejection from the BFW preheaters, cooler and boiler across the entire proposed
system are presented in Fig. 16 and Fig. 17. While work consumption and work production are
shown in Fig. 18. Finally, the overall system performance is found based on the proposed exergy
efficiency definition as illustrated in Table 11. The ammonia production performance is found to
be 71 % while taking into consideration the steam recovery, the second exergy efficiency of the
integrated system is found to be 62 % and taking into consideration excess gas and transient
condition the system attained an exergy efficiency of 73%., respectively.

12
5 Conclusions

After conducting a comprehensive exergy analysis and assessment of the newly developed
integrated ammonia production system, the following findings are concluded from this research:
• The small exergy consumption and losses across the reformer unit are related to the flux and
heat transfer across drastic changes in temperature due to irreversibilities.
• The integrated expander unit in the ammonia synthesizer shows great potential for improved
ammonia conversion where exergy analysis measures the performance of the process close to
ideal.
• The exergy losses and consumption of the CO shift conversion, CO2 removal, and
methanation are insignificant because of the low-quality exergy.
• The overall exergy efficiency of the system is 71% and turbine exergy efficiency of 92 % for
optimal operating conditions with a total work rate of 12631 kW and ammonia conversion
rate of 0.38.
In summary, the results shown in this study can be further investigated and is useful for
researches in the field of optimization of industrial ammonia unit.

13
Acknowledgment
The authors acknowledge the support by the Natural Sciences and Engineering Research Council
of Canada (NSERC).

Nomenclature
Ac catalyst activity
C concentration (M)


energy (kJ)


energy rate (kW)
 exergy rate (kW)
Ex exergy (kJ)
 specific exergy (kJ/kg)
F fugacity
ℎ specific enthalpy (kJ/kg)
K kinetic constant
KHT High temperature kinetic constant
KLT standard LT catalyst activity
KLT Low temperature kinetic constant
LHV lower heating value (kJ)
 mass (kg)


mass flow rate (kg/s)
ƒ molar flow rate (mol/s)
P pressure (Pa or kPa)
R rate of conversion
2 universal gas constant (8.314 kJ/kmol K)
S number of moles of methane

R
s specific entropy (kJ/kgK)
entropy rate (kW/K)
T temperature
 temperature (oC)
t time (seconds)

¨
Tref reference temperature
volume flow rate (m3/s)
#

work (kJ)
# work rate (W or kW)
Y mole fraction
Xi mole fraction of component i
Greek letters
© transfer coefficient
δ Nernst diffusion layer thickness
∆ change
ª efficiency
ρ density (kg/m3)
π Pi number
ξ reaction rate
14
Acronyms
NH3 ammonia
H2 hydrogen
N2 nitrogen
CH4 methane
Ar argon
CO carbon monoxide
CO2 carbon dioxide
H2O water
O2 oxygen
C2H6 ethane

15
References

[1] Alexandratos N, Bruinsma J. World agriculture: towards 2015/2030: an FAO perspective.


vol. 20. Rome: 2003. doi:10.1016/S0264-8377(03)00047-4.
[2] Kirova-Yordanova à Z, Zlatarov A, Str Y. Exergy analysis of industrial ammonia
synthesis. Energy 2004;29:2373–84. doi:10.1016/j.energy.2004.03.036.
[3] Greeff IL, Visser JA, Ptasinski KJ, Janssen FJJG. Integration of a turbine expander with
an exothermic reactor loop—Flow sheet development and application to ammonia
production. Energy 2003;28:1495–509. doi:10.1016/S0360-5442(03)00122-1.
[4] Flórez-Orrego D, de Oliveira Junior S. Modeling and optimization of an industrial
ammonia synthesis unit: An exergy approach. Energy 2017;137:234–50.
doi:10.1016/J.ENERGY.2017.06.157.
[5] Flórez-Orrego D, de Oliveira Junior S. Exergy assessment of single and dual pressure
industrial ammonia synthesis units. Energy 2017;141:2540–58.
doi:10.1016/J.ENERGY.2017.06.139.
[6] Khademi MH, Sabbaghi RS. Comparison between three types of ammonia synthesis
reactor configurations in terms of cooling methods. Chem Eng Res Des 2017;128:306–17.
doi:10.1016/J.CHERD.2017.10.021.
[7] Sahafzadeh M, Ataei A, Tahouni N, Panjeshahi MH. Integration of a gas turbine with an
ammonia process for improving energy efficiency. Appl Therm Eng 2013;58:594–604.
doi:10.1016/J.APPLTHERMALENG.2013.05.006.
[8] Aziz M, Oda T, Morihara A, Kashiwagi T. Combined nitrogen production, ammonia
synthesis, and power generation for efficient hydrogen storage. Energy Procedia
2017;143:674–9. doi:10.1016/J.EGYPRO.2017.12.745.
[9] Rosen MA. Thermodynamic investigation of hydrogen production by steam-methane
reforming. Int J Hydrogen Energy 1991;16:207–17. doi:10.1016/0360-3199(91)90003-2.
[10] Simpson AP, Lutz AE. Exergy analysis of hydrogen production via steam methane
reforming. Int J Hydrogen Energy 2007;32:4811–20.
doi:10.1016/J.IJHYDENE.2007.08.025.
[11] Dincer I, Rosen MA. Chemical Exergy. Exergy, Elsevier; 2013, p. 31–49.
doi:10.1016/B978-0-08-097089-9.00003-6.
[12] Alatiqi IM, Meziou AM, Gasmelseed GA. Static and Dynamic Simulation of Steam
Methane Reformers, 1989, p. 535–50. doi:10.1016/S0167-2991(08)61088-5.
[13] Sehested J, Dahl S, Jacobsen J, Rostrup-Nielsen JR. Methanation of CO over
Nickel: Mechanism and Kinetics at High H 2 /CO Ratios †. J Phys Chem B
2005;109:2432–8. doi:10.1021/jp040239s.
[14] Lim Y, Kim J, Jung J, Lee CS, Han C. Modeling and Simulation of CO 2 Capture Process
for Coal-based Power Plant using Amine Solvent in South Korea Selection and/or peer-
review under responsibility of GHGT. Youngsub Lim Al / Energy Procedia
2013;37:1855–62. doi:10.1016/j.egypro.2013.06.065.
[15] Yadav R, Rinker RG. Steady‐state methanation kinetics over a Ni/Al2O3 catalyst. Can J
Chem Eng 1993;71:202–8. doi:10.1002/cjce.5450710206.
[16] Singh CPP, Saraf DN. Simulation of Ammonia Synthesis Reactors. Ind Eng Chem
Process Des Dev 1979;18:364–70. doi:10.1021/i260071a002.
[17] Gillespie LJ, Beattie JA. The thermodynamic treatment of chemical equilibria in systems
composed of real gases. I. An approximate equation for the mass action function applied
16
to the existing data on the haber equilibrium. Phys Rev 1930;36:743–53.
doi:10.1103/PhysRev.36.743.
[18] Ukpaka, Izonowei. International Scientific Orgnization. Chem Int 2017;3:46–57.
[19] Dincer I, Zamfirescu C. Energy, Environment, and Sustainable Development. Adv. Power
Gener. Syst., Elsevier; 2014, p. 55–93. doi:10.1016/B978-0-12-383860-5.00002-X.
[20] Sorin M, Lambert J, Paris J. Exergy Flows Analysis in Chemical Reactors. Chem Eng Res
Des 1998;76:389–95. doi:10.1205/026387698524811.
[21] Bargigli S, Raugei M, Ulgiati S. Comparison of thermodynamic and environmental
indexes of natural gas, syngas and hydrogen production processes. Energy 2004;29:2145–
59. doi:10.1016/J.ENERGY.2004.03.018.
[22] Li Y, Shang K, Lu N. IOP Conference Series: Earth and Environmental Science Aqueous
ammonia process for CO2 capture Related content Oxidation of ammonium sulfite in
aqueous solutions using ozone technology. IOP Conf Ser Earth Environ Sci
2009;6:172017. doi:10.1088/1755-1307/6/7/172017.
[23] Pinsent BRW, Roughton FJW. The kinetics of combination of carbon dioxide with water
and hydroxide ions. Trans Faraday Soc 1951;47:263. doi:10.1039/tf9514700263.

17
Table 1 Input streams for the integrated energy system.

Stream Name A-1 BFW2 BFW-1 S1


Temperature 33.00 20.00 130.00 502.55
Pressure (kPa) 294.20 200.00 12748.60 3530.39
Mole flow (Kmol/s) 0.66 2.78 0.99 1.99
H2 (Kmol/s) 0.00 0.00 0.00 0.02
N2 (Kmol/s) 0.51 0.00 0.00 0.01
CH4 (Kmol/s) 0.00 0.00 0.00 0.32
Ar (Kmol/s) 0.01 0.00 0.00 0.08
NH3 (Kmol/s) 0.00 0.00 0.00 0.00
H2O (Kmol/s) 0.00 2.78 0.99 1.57
O2 (Kmol/s) 0.14 0.00 0.00 0.00

Table 2 Standard chemical exergy for selected substances.

Substance Chemical exergy (kJ/kmol k)


CO2 19870
CO 274710
H2 236090
N2 720
CH4 831200
Ar 12000
NH3 337900
H2O 9500
O2 3970

Table 3 The kinetic parameters for reactions are derived from literature [23].
R. No. K E (cal/mol)
13
iv 4.32x10 13249
17
v 2.38x10 29451
11
vi 1.35x10 11585
20
vii 4.75x10 16529

Table 4 The values for A, B, and C parameters [23].


Reaction A B C
NH4HCO3 salt -914.00821 38648.2117 136.174996

18
CH4

Fig. 1 The schematic representation of the integrated energy system.

19
Power generation

Subsystem A Subsystem D

Subsystem B Subsystem C

Process
steam and
natural gas

Subsystem E

Ammonia loop

Fig. 2 Aspen plus model of the integrated energy system. The dashed lines represent the border for each subsystem /proces

20
A-3(IN) R1 2430.92 2946.04
193.28 0.24
434.35 SREF-R 691.38
SREF-S 5196.46 5067.43

R5 R6 S2(OUT)
H2 mole flow rate (kmol/hr)

CH4 mole flow rate (kmol/hr)

CO mole flow rate (kmol/hr)


3423.57
H2O mole flow rate (kmol/hr)
193.28
434.35
R4 4203.79
62.86
1143.48
0.00 PREF-S PREF-T
5664.13

S1(IN) R2 62.86
1142.05
0.00
5666.98

R3

Fig. 3 Subsystem A represents the steam reforming process. S1 and A-3 are the inputs and S2 is the
output. PREF-S and SREF-S are RSTOICH reactors for combustion while PREF-T and SREF-R are
Rplug reactors and represent the number of firing zones.

Stream Name Units R1 R2 R3 R4 R5 R6

Temperature (°°C) C 492.00 502.55 505.62 821.27 1318.22 1186.02


Pressure (kPa) kPa 3236.19 3530.39 3340.39 3068.93 3068.93 2872.80
Enthalpy Flow (kW) kW 9205.38 -369806 -369806 -286375 -277170 -281386
Mole Flows
kmol/s 0.66 1.99 1.99 2.52 3.04 3.15
(kmol/s)
CO2 (kmol/s) kmol/s 0.00 0.00 0.00 0.14 0.14 0.13
CO (kmol/s) kmol/s 0.00 0.00 0.00 0.12 0.12 0.19
H2 (kmol/s) kmol/s 0.00 0.02 0.02 0.95 0.68 0.82
N2 (kmol/s) kmol/s 0.51 0.01 0.01 0.01 0.52 0.52
CH4 (kmol/s) kmol/s 0.00 0.32 0.32 0.05 0.05 0.00
Ar (kmol/s) kmol/s 0.01 0.08 0.08 0.08 0.08 0.08
H2O (kmol/s) kmol/s 0.00 1.57 1.57 1.17 1.44 1.41
O2 (kmol/s) kmol/s 0.14 0.00 0.00 0.00 0.00 0.00
Physical exergy rate
kW 9403.58 43618.7 43554.6 64466.9 125874 112556
(kW)
Chemical exergy
kW 70.48 280675.65 280369.14 309395.73 244328.93 252303.72
rate (kW)
Total exergy rate kW 9474.06 324294.35 323923.74 373862.63 370202.93 364859.72

21
Table 5 Subsystem A results. R1 to R6 are the streams input and output of the reformer process

Fig. 4 The exergy loss and consumption across the reactors in the reformer unit/ Subsystem A.

Fig. 5 The enthalpy of formation for the reaction taking place in the reformer unit/ subsystem A versus
temperature. The negative enthalpy flow rate shows that the reaction across the unit is endothermic.
22
Fig. 6 Subsystem B represents the carbon monoxide removal process. The high-temperature shift and
low-temperature shift reactors are HT-SHIFT and LT-SHIFT respectively.

Table 6 Subsystem B results. CO-1 to CO-11 are the streams input and output of the carbon monoxide
removal process.

23
Stream
CO-1 CO-2 CO-3 CO-4 CO-5 CO-6 CO-7 CO-8 CO-9 CO-10 CO-11
Name
Temperature
380 435 435 435 340 340 340 130 300 210 217
(°C)
Pressure
2872 2804. 2804 2804 2804 2804 2804 12748 12650 2779 2651
(kPa)
Mole Flows
3.147 3.147 2.989 0.157 2.989 0.157 3.147 0.992 0.992 3.147 3.147
(kmol/s)
CO2 (kmol/s) 0.126 0.287 0.273 0.014 0.273 0.014 0.287 0.000 0.000 0.287 0.316

CO (kmol/s) 0.192 0.031 0.029 0.002 0.029 0.002 0.031 0.000 0.000 0.031 0.0083

H2 (kmol/s) 0.818 0.980 0.931 0.049 0.931 0.049 0.980 0.000 0.000 0.980 1.002

N2 (kmol/s) 0.522 0.522 0.495 0.026 0.495 0.026 0.522 0.000 0.000 0.522 0.522

Ar (kmol/s) 0.082 0.082 0.077 0.004 0.077 0.004 0.082 0.000 0.000 0.082 0.082

NH3 (kmol/s) 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000

H2O (kmol/s) 1.408 1.246 1.184 0.062 1.184 0.062 1.246 0.992 0.992 1.246 1.223
Physical
exergy rate 46055.9 48156 45748 2407 40309 2121 42431 1351.53 6785 35996 35861
(kW)
Chemical
exergy rate 252303 247868 235475 12393 235475 12393 247868 9421.63 9421.6 247868 247260.5
(kW)
Total exergy
298359 296024 281223 14801 275785 14515 290300 10773 16207 283864 283121.8
rate (kW)

24
Fig. 7 The exergy loss and consumption across the CO removal process.

0.000
1627.186
9222.19

CO2R-1

1137.211 0.381

0.000 15.853

4382.75 8.93

S5(IN) CO2R-2 ABSORBER CO2R-3 S6(OUT)

0.288
CO2 mole flow rate (kmol/hr)
2.829

NH3 mole flow rate (kmol/hr) 40.59

H2O mole flow rate (kmol/hr) CO2R-5


0.423
50.523
12810.71

CO2R-4
STRIPPER
0.135
47.694
12770.14

CO2R-6

Fig. 8 Subsystem C represents the carbon dioxide stripping process. The process consists of a 2 stage;
absorber reactor and stripper reactor.

25
Table 7 Subsystem C results. CO2R-1 to CO2R-5 are the streams input and output of the CO2 capturing
process.

Stream Name CO2R-1 CO2R-2 CO2R-3 CO2R-4 CO2R-5 C02R-6


Temperature (°C) 30 40 30 30 60 60.00
Pressure (kPa) 2651.66 2651.66 2651.66 2651.66 101.33 101.33
Enthalpy Flow (kW) -756638 -473650 -833.42 -126175 -2751 -1079980
Mole Flows (kmol/s) 3.014 3.147 1.565 4.192 0.065 4.13
CO2 (kmol/s) 0.000 0.316 0.000 0.000 0.000 0.00
CO (kmol/s) 0.000 0.002 0.002 0.000 0.000 0.00
H2 (kmol/s) 0.000 1.009 1.009 0.000 0.000 0.00
N2 (kmol/s) 0.000 0.522 0.480 0.042 0.040 0.00
CH4 (kmol/s) 0.000 0.000 0.000 0.000 0.000 0.00
Ar (kmol/s) 0.000 0.082 0.068 0.014 0.013 0.00
NH3 (kmol/s) 0.452 0.000 0.004 0.014 0.001 0.01
H2O (kmol/s) 2.562 1.217 0.002 3.559 0.011 3.55
NH4+ (kmol/s) 0.00 0.00 0.00 0.25 0.00 0.25
NH2COO- (kmol/s) 0.00 0.00 0.00 0.10 0.00 0.10
HCO3- (kmol/s) 0.00 0.00 0.00 0.11 0.00 0.11
CO3- (kmol/s) 0.00 0.00 0.00 0.02 0.00 0.02
NH4HCO3S (kmol/s) 0.00 0.00 0.00 0.09 0.00 0.09
Physical exergy rate (kW) 424.74 15869.20 12704.70 2337.82 30.63 8480.00
Chemical exergy rate (kW) 173907.8 247260.28 238100.57 36497.26 467.98 36617.25
Total exergy rate (kW) 174332.54 263129.48 250805.27 38835.08 498.61 45097.25

Fig. 9 Exergy destruction of subsystem C/ CO2 capturing process.

26
3549.66
6.50
0.00
3630.71 3611.05
0.14 R-M1 6.56 M-3 S8(OUT)
8.93 15.74

SPLIT-1
S7(IN) M-1 M-2
61.39
H2 mole flow rate (kmol/hr) 0.07
CH4 mole flow rate (kmol/hr) 15.74
H2O mole flow rate (kmol/hr)

M-4

Fig. 10 Subsystem D represents the methanation process. R-M1 is modeled as an Rplug reactor.

Table 8 Subsystem D results. M-1 to M-4 are the streams input and output of the methanation process.

Stream Name M-1 M-2 M-3 M-4

Temperature (°C) 280.00 280.00 280.00 280.00

Pressure (kPa) 2651.66 2651.66 2651.66 2651.66

Enthalpy Flow (kW) 10569.90 10187.90 11199.70 -1013.52

Mole Flows (kmol/s) 1.56 1.56 1.53 0.03

H2 (kmol/s) 1.01 1.00 0.99 0.02

N2 (kmol/s) 0.48 0.48 0.48 0.00

CH4 (kmol/s) 0.00 0.00 0.00 0.00

Ar (kmol/s) 0.07 0.07 0.07 0.00


Physical exergy rate
15944.30 15910.10 15586.10 334.84
(kW)
Chemical exergy rate
238100.58 237957.57 232435.33 5484.58
(kW)
Total exergy rate (kW) 254044.88 253867.67 248021.43 5819.42

27
Fig. 11 Subsystem E represents the ammonia synthesis process. The unit consists of 3 Rplug reactors; R1-A, R1-B and R1-C, and 1 convertor
reactor R1-D.

28
Table 9 Subsystem D results. SYN-1 to SYN-15 are the streams input and output of the ammonia synthesis process.

Stream Name SYN- SYN- SYN- SYN- SYN- SYN- SYN- SYN- SYN- SYN- SYN- SYN- SYN- SYN- SYN-
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Pressure 3000 3000 3000 3000 3000 3000 3000 2980 2980 2960 2960 2930 2930 2920 3000
(kPa) 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0
Temperature 180 180 179 289 386 465 452 532 456 489 395 431 321 391 179

Enthalpy Flow 7297 9833 1678 3402 4889 6104 6138 6138 4923 4923 3436 3436 1713 1713 343
(kW) 7 2 0 1 4 4 2 2 5 5 1 1
Mole Flows 1.53 3.73 5.16 5.16 5.16 5.16 5.26 5.03 5.03 4.94 4.94 4.83 4.83 4.64 0.11
(kmol/s)
H2 0.99 0.62 1.58 1.58 1.58 1.58 1.61 1.26 1.26 1.12 1.12 0.97 0.97 0.67 0.03
(kmol/s)
N2 0.48 2.12 2.54 2.54 2.54 2.54 2.60 2.48 2.48 2.43 2.43 2.38 2.38 2.28 0.05
(kmol/s)
Ar 0.07 0.87 0.91 0.91 0.91 0.91 0.93 0.93 0.93 0.93 0.93 0.93 0.93 0.93 0.02
(kmol/s)
NH3 0.00 0.10 0.10 0.10 0.10 0.10 0.10 0.33 0.33 0.42 0.42 0.53 0.53 0.73 0.00
(kmol/s)
Physical exergy rate 2369 5722 7932 8638 9397 1009 1024 1059 9853 9988 9118 9241 8308 8525 1619
(kW) 9 3 2 3 3 30 53 97 3 0 0 6 0 4
Chemical exergy 2324 2013 4235 4235 4235 4235 4322 4278 4278 4262 4262 4246 4246 4218 8645
rate 35 24 86 82 81 81 25 77 77 95 95 75 73 45
(kW)
Total exergy rate 2561 2585 5029 5099 5175 5245 5346 5338 5264 5261 5174 5170 5077 5070 1026
(kW) 33 48 08 65 54 11 78 74 10 75 75 90 53 99 3

29
Fig. 12 The enthalpy of reaction versus temperature in the ammonia synthesis sub-process.

Fig. 13 The conversion mole fraction in the ammonia synthesis process. While the equilibrium line shows
the optimal conversion rate for ammonia synthesis unit.

30
Fig. 14 The exergy loss and consumption across the integrated system sub-processes.

Fig. 15 The exergy efficiency across the integrated energy system.


31
Fig. 16 The heat consumptions of the heat exchangers across the integrated energy system.

Fig. 17 The heat rejection of the boilers across the integrated energy system.

Fig. 18 Power consumption and production across the integrated energy system.

32
Table 10 Subsystem Ammonia loop results. S9 to PURGE are the streams input and output of the ammonia loop.

Stream Name S9 S10 S11 S12 S13 S14 S15 LIQ-NH3 PURGE

Temperature (°°C) 180.00 391.31 199.66 -33.00 -33.00 209.54 180.00 -33.00 -33.00
Pressure (kPa) 30000.00 29200.00 4800.00 4800.00 4800.00 30000.00 30000.00 4800.00 4800.00
Enthalpy Flow
(kW) 7296.75 16824.70 -12409.70 -59560.50 -14401.10 13025.40 9527.93 -44075.40 -1083.96
Mole Flows
(kmol/s) 1.53 4.70 4.70 4.70 3.79 3.79 3.79 0.62 0.29
H2 (kmol/s) 0.99 0.67 0.67 0.67 0.63 0.63 0.63 0.00 0.05
N2 (kmol/s) 0.48 2.31 2.31 2.31 2.15 2.15 2.15 0.00 0.16
CH4 (kmol/s) 0.00 0.03 0.03 0.03 0.02 0.02 0.02 0.00 0.00
Ar (kmol/s) 0.07 0.95 0.95 0.95 0.88 0.88 0.88 0.00 0.07
NH3 (kmol/s) 0.00 0.74 0.74 0.74 0.11 0.11 0.11 0.62 0.01
Physical exergy 23698.7 86118.2 50091.3 47353.2 37062.7 59396.8 58128.7 3807.63 2789.67
rate (kW)
Chemical exergy 232435.331 427258 427258.008 427258.008 206706.897 206706.897 206706.897 208693.235 15558.6024
rate (kW)
Total exergy rate 256134.031 513376.2 477349.308 474611.208 243769.597 266103.697 264835.597 212500.865 18348.2724
(kW)

33
Table 11 Exergy efficiency definitions

Equation Exergy efficiency Results

­®¯° ±²®¯° + ³´±µ


«¬ =
­¶·´¸¹¶ ±²¶·´¸¹¶ + ­¶µ±¹­ ±²¶µ±¹­ + ­¹º» ±²¹º»
(I) Overall 71%

­»±½¾¿±»±À ¶µ±¹­ ±²ÁÂÃÄÅÂÁÂÆ ¶µ±¹­


«¼ =
­¶·´¸¹¶ ±²¶·´¸¹¶ + ­¹º» ±²¹º» − ­®¯° ±²®¯° − ³´±µ
(II) Steam 62%
Recovery

³´±µ
«° = 73%
­¶·´¸¹¶ ±²¶·´¸¹¶ + ­¶µ±¹­ ±²¶µ±¹­ " − ­Dz½±¶¶ ¸¹¶ ±²Ç²½±¶¶ ¸¹¶ − ­®¯° ±²®¯°
(III)
Unreacted gas
transit

34
Highlights
• Integrated ammonia synthesis loop with power production
• Energy platform for power, steam and ammonia production
• Performance assessment through exergy efficiencies
Conflict of Interest and Authorship Conformation Form

Please check the following as appropriate:

o All authors have participated in (a) conception and design, or analysis and
interpretation of the data; (b) drafting the article or revising it critically for
important intellectual content; and (c) approval of the final version.
Yes
o This manuscript has not been submitted to, nor is under review at, another
journal or other publishing venue.
Yes
o The authors have no affiliation with any organization with a direct or indirect
financial interest in the subject matter discussed in the manuscript
Yes
o The following authors have affiliations with organizations with direct or
indirect financial interest in the subject matter discussed in the manuscript:

You might also like